Skip to main content
Log in

Initial investigation into the complementary use of black box and physics-based techniques in rotorcraft system identification

  • Original Paper
  • Published:
CEAS Aeronautical Journal Aims and scope Submit manuscript

Abstract

Accurate linear helicopter models are needed for control system development and simulation and can be determined by system identification when appropriate test data are available. Standard methods for rotorcraft system identification are the frequency domain maximum likelihood method and the frequency response method that are used to derive physics-based linear state-space models. Also the optimized predictor-based subspace identification method (PBSIDopt), a time domain system identification method that yields linear black box state-space models, has been successfully applied to rotorcraft data. As both methods have their respective strengths and weaknesses, it was tried to combine both techniques. The paper demonstrates the successful complementary use of physics-based frequency domain methods and the black box PBSIDopt method in the areas of database requirements, accuracy metrics, and model structure development using flight test data of DLR’s ACT/FHS research rotorcraft.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8

Similar content being viewed by others

Abbreviations

\(a_x\), \(a_y\), \(a_z\) :

Longitudinal, lateral, and vertical acceleration (m/s\(^2\))

\(\varvec{A}\), \(\varvec{B}\), \(\varvec{C}\), \(\varvec{D}\) :

State-space matrices (continuous time)

\(\mathrm {CR}_j\) :

Cramer–Rao bound of the jth parameter

\(\varvec{\mathcal {F}}\) :

Fischer information matrix

J :

Cost function

L, M, N :

Moment derivatives

n :

Model order

\(n_y\) :

Number of model outputs

p, q, r :

Roll, pitch and yaw rates (rad/s)

\(\varvec{R}\) :

Measurement noise covariance matrix

u, v, w :

Airspeed components (aircraft fixed) (m/s)

\(\varvec{u}\), \(\varvec{x}\), \(\varvec{y}\) :

Input, state, and output vectors

X, Y, Z :

Force derivatives

\(\delta _{\mathrm{{lon}}}\), \(\delta _{\mathrm{{lat}}}\) :

Longitudinal and lateral cyclic inputs (%)

\(\delta _{col}\), \(\delta _{\mathrm{{ped}}}\) :

Collective and pedal inputs (%)

\(\phi\), \(\theta\) :

Roll and pitch attitude angles (rad)

\(\varvec{\varTheta }\) :

Unknown model parameters

ACT/FHS:

Active control technology/flying helicopter simulator

CR:

Cramer–Rao

DLR:

German Aerospace Center

FR:

Frequency response

ML:

Maximum likelihood

PBSIDopt:

Optimized predictor-based subspace identification (method)

References

  1. Hamel, P., Jategaonkar, R.: Evolution of flight vehicle system identification. J. Aircraft 33(1), 9–28 (1996)

    Article  Google Scholar 

  2. Tischler, M.B., Remple, R.K.: Aircraft and Rotorcraft System Identification: Engineering Methods with Flight-Test Examples, 2nd edn. American Institute of Aeronautics and Astronautics Inc, Reston (2012)

    Book  Google Scholar 

  3. Chiuso, A.: The role of vector autoregressive modeling in predictor-based subspace identification. Automatica 43(6), 1034–1048 (2007). https://doi.org/10.1016/j.automatica.2006.12.009

    Article  MathSciNet  MATH  Google Scholar 

  4. Chiuso, A.: On the asymptotic properties of closed-loop CCA-type subspace algorithms: equivalence results and role of the future horizon. IEEE Trans. Autom. Control 55(3), 634–649 (2010). https://doi.org/10.1109/TAC.2009.2039239

    Article  MathSciNet  MATH  Google Scholar 

  5. Bergamasco, M., Lovera, M.: Continuous-time predictor-based subspace identification for helicopter dynamics. In: 37th European Rotorcraft Forum. Gallarate, Italy (2011)

  6. Bergamasco, M., Cortigiani, N., Del Gobbo, D., Panza, S., Lovera, M.: The role of black-box models in rotorcraft attitude control. In: 43rd European Rotorcraft Forum. Milano, Italy (2017)

  7. Greiser, S., Seher-Weiss, S.: A contribution to the development of a full flight envelope quasi-nonlinear helicopter simulation. CEAS Aeronaut. J. 5(1), 53–66 (2014). https://doi.org/10.1007/s13272-013-0090-z

    Article  Google Scholar 

  8. Seher-Weiss, S.: Comparing different approaches for modeling the vertical motion of the EC 135. CEAS Aeronaut. J. 6(3), 395–406 (2015). https://doi.org/10.1007/s13272-015-0150-7

    Article  Google Scholar 

  9. Seher-Weiss, S., von Gruenhagen, W.: Development of EC 135 turbulence models via system identification. Aerosp. Sci. Technol. 23(1), 43–52 (2012). https://doi.org/10.1016/j.ast.2011.09.008

    Article  Google Scholar 

  10. Seher-Weiss, S., von Grünhagen, W.: EC 135 system identification for model following control and turbulence modeling. In: 1st CEAS European Air and Space Conference, pp. 2439–2447. Berlin, Germany (2007)

  11. Seher-Weiss, S., von Grünhagen, W.: Comparing explicit and implicit modeling of rotor flapping dynamics for the EC 135. CEAS Aeronaut. J. 5(3), 319–332 (2014). https://doi.org/10.1007/s13272-014-0109-0

    Article  Google Scholar 

  12. Wartmann, J., Seher-Weiss, S.: Application of the predictor-based subspace identification method to rotorcraft system identification. In: 39th European Rotorcraft Forum. Moscow, Russia (2013)

  13. Wartmann, J.: ACT/FHS system identification including engine torque and main rotor speed using the PBSIDopt method. In: 41st European Rotorcraft Forum. Munich, Germany (2015)

  14. Wartmann, J., Greiser, S.: Identification and selection of rotorcraft candidate models to predict handling qualities and dynamic stability. In: 44th European Rotorcraft Forum. Delft, The Netherlands (2018)

  15. Wartmann, J.: Closed-loop rotorcraft system identification using generalized binary noise. In: AHS 73rd Annual Forum. Fort Worth, TX (2017)

  16. Chiuso, A.: On the relation between CCA and predictor-based subspace identification. IEEE Trans. Autom. Control 52(10), 1795–1812 (2007). https://doi.org/10.1109/TAC.2007.906159

    Article  MathSciNet  MATH  Google Scholar 

  17. van der Veen, G., van Wingerden, J.W., Bergamasco, M., Lovera, M., Verhaegen, M.: Closed-loop subspace identification methods: an overview. IET Control Theory Appl. 7(10), 1339–1358 (2013). https://doi.org/10.1049/iet-cta.2012.0653

    Article  MathSciNet  Google Scholar 

  18. Fragniére, B., Wartmann, J.: Local polynomial method frequency-response calculation for rotorcraft applications. In: AHS 71st Annual Forum. Virginia Beach, VA (2015)

  19. Seher-Weiß, S.: ACT/FHS system identification including rotor and engine dynamics. In: AHS 73rd Annual Forum. Fort Worth, TX (2017)

  20. Maine, R.E., Iliff, K.W.: Theory and practice of estimating the accuracy of dynamic flight-determined coefficients. Tech. Rep. RP 1077, NASA (1981)

  21. van Wingerden, J.W.: The asymptotic variance of the PBSIDopt algorithm. IFAC Proc. Vol. 45(16), 1167–1172 (2012). https://doi.org/10.3182/20120711-3-BE-2027.00228

    Article  Google Scholar 

  22. Greiser, S., von Gruenhagen, W.: Improving system identification results combining a physics-based stitched model with transfer function models obtained through inverse simulation. In: American Helicopter Society 72nd Annual Forum. West Palm Beach, Florida, USA (2016)

  23. Bauer, F., Lukas, M.A.: Comparing parameter choice methods for regularization of ill-posed problems. Math. Comput. Simul. 81(9), 1795–1841 (2011)

    Article  MathSciNet  Google Scholar 

  24. Ljung, L., McKelvey, T.: Subspace identification from closed loop data. Signal Process. 52(2), 209–215 (1996). https://doi.org/10.1016/0165-1684(96)00054-0

    Article  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Susanne Seher-Weiß.

Appendix: Applied identification methods

Appendix: Applied identification methods

\(\varvec{A}_d\), \(\varvec{B}_d\), \(\varvec{C}_d\), \(\varvec{D}_d\):

Discrete-time state-space matrices

\(\varvec{A}_K\), \(\varvec{B}_K\):

Predictor form state-space matrices

\(\varvec{e}_k\), \(\varvec{u}_k\), \(\varvec{x}_k\), \(\varvec{y}_k\):

Discrete-time innovation, input, state, and output vectors at kth time step

\(\varvec{E}\), \(\varvec{U}\), \(\varvec{X}\), \(\varvec{Y}\):

Data matrices for system innovation, input, state, and output

f, p:

Future and past window length

\(\varvec{K}\) :

Kalman gain matrix

\(\varvec{\mathcal {K}}^{(p)}\) :

Extended controllability matrix

\(M_f\), \(M_n\), \(M_p\):

Sets for f, n and p

\(n_u\) :

Number of model inputs

N :

Number of data points

\(\varvec{\mathcal {O}}^{(f)}\) :

Extended observability matrix

s :

Laplace variable (1/s)

\(\varvec{\mathcal {S}}\) :

Diagonal singular values matrix

\(\varvec{T}\) :

Frequency response matrix

\(w_{\mathrm{{ap}}}\) :

Relative weighting amplitude/phase errors

\(w_{\gamma }\) :

Coherence weighting

\(\varvec{y}_{\mathrm{{m}},k}\) :

Measured output (index m)

\(\varvec{z}_k\) :

Merged input–output vector at kth time step

\(\varvec{Z}\) :

Data matrices for merged input–outputs (used with indexes)

\(\gamma _{uy}^2\) :

Coherence between u and y

\(\lambda\) :

Regularization parameter

\(\omega\) :

Angular frequency (rad/s)

\(\sigma (\ldots )\) :

Standard deviation

\(\tau\) :

Time delay (s)

\(\measuredangle\) :

Phase angle (\(^{\circ }\))

\(|\ldots |_{\text{dB}}\) :

Amplitude (dB)

1.1 ML output error method

The system to be identified is assumed to be described by a linear state-space model

$$\begin{aligned} \begin{aligned} \dot{\varvec{x}}(t)&= \varvec{A}(\varvec{\varTheta }) \varvec{x}(t) + \varvec{B}(\varvec{\varTheta }) \varvec{u}(t)\\ \varvec{y} (t)&= \varvec{C}(\varvec{\varTheta }) \varvec{x}(t) + \varvec{D}(\varvec{\varTheta }) \varvec{u}(t), \end{aligned} \end{aligned}$$
(6)

where \(\varvec{x}\) denotes the state vector, \(\varvec{u}\) the input vector and \(\varvec{y}\) the output vector. The system matrices \(\varvec{A}\), \(\varvec{B}\), \(\varvec{C}\) and \(\varvec{D}\) contain the unknown model parameters \(\varvec{\varTheta }\). Measurements \(\varvec{z}\) of the outputs exist for N discrete time points \(t_k\)

$$\begin{aligned} \varvec{z}_k = \varvec{y}(t_k) + \varvec{v}(t_k) \ , \ \ \ k=1, \dots , N. \end{aligned}$$
(7)

The measurement noise \(\varvec{v}\) is assumed to be characterized by Gaussian white noise with covariance matrix \(\varvec{R}\).

The ML estimates of the unknown parameters \(\varvec{\varTheta }\) and of the measurement noise covariance matrix \(\varvec{R}\) are obtained by minimizing the cost function

$$\begin{aligned} \begin{aligned} J(\varvec{\varTheta },\varvec{R})&= \frac{1}{2} \sum _{k=1}^{N}{\left[ \varvec{z}(t_k)-\varvec{y}(t_k) \right] ^T \varvec{R}^{-1}} \\&\times \left[ \varvec{z}(t_k)-\varvec{y}(t_k) \right] + \frac{N}{2} \ln {\left( \det {(\varvec{R})}\right) } . \end{aligned} \end{aligned}$$
(8)

If the measurement error covariance matrix \(\varvec{R}\) is unknown, as it is usually the case, the optimization of Eq. (8) is carried out in two steps. In the first step, it can be shown that for any given value of \(\varvec{\varTheta }\), the ML estimate of \(\varvec{R}\) is given by

$$\begin{aligned} \varvec{R} = \frac{1}{N} \sum _{k=1}^{N}{\left[ \varvec{z}(t_k)-\varvec{y}(t_k) \right] \left[ \varvec{z}(t_k)-\varvec{y}(t_k) \right] ^T } \end{aligned}$$
(9)

which means that the output error covariance matrix is the most plausible estimate for \(\varvec{R}\).

Thus, the variable part of the cost function reduces to

$$\begin{aligned} J(\varvec{\varTheta }) = \ln {\left( \det {(\varvec{R})}\right) }. \end{aligned}$$
(10)

If the covariance matrix \(\varvec{R}\) is assumed to be a diagonal matrix, the cost function reduces to the product of the output error variances of all output variables

$$\begin{aligned} J(\varvec{\varTheta }) = \prod _{j=1}^{n_y} \left( \frac{1}{N} \sum _{k=1}^{N}{\left[ \varvec{z}_j(t_k)-\varvec{y}_j(t_k) \right] ^2} \right) . \end{aligned}$$
(11)

Frequency domain variant

The discretely sampled time-dependent variable

$$\begin{aligned} \varvec{x}_k = \varvec{x}(k\varDelta t) \ , \ \ \ k=0, \ldots , N-1 \end{aligned}$$
(12)

with the sampling time interval \(\varDelta t\) is transformed to a frequency-dependent variable using the Fourier transform

$$\begin{aligned} \begin{aligned} \varvec{x}(\omega _k)&= \frac{1}{N} \sum _{k=0}^{N-1} \varvec{x}_k \text {e}^{-\text {i}\omega _k k \varDelta t} \\ \omega _k&=k \cdot 2\pi / t_N \ \ \text {with} \ \ t_N =(N-1)\varDelta t \ . \end{aligned} \end{aligned}$$
(13)

Transforming the variables \(\dot{\varvec{x}}\), \(\varvec{x}\), \(\varvec{u}\), \(\varvec{y}\) of the linear model from Eq. (6) to the frequency domain leads to the following model equations in the frequency domain

$$\begin{aligned} \begin{aligned} \text {i} \omega \varvec{x}(\omega )&= \varvec{A}(\varvec{\varTheta }) \varvec{x}(\omega ) + \varvec{B}(\varvec{\varTheta }) \varvec{u}(\omega ) \\ \varvec{y}(\omega )&= \varvec{C}(\varvec{\varTheta }) \varvec{x}(\omega ) + \varvec{D}(\varvec{\varTheta }) \varvec{u}(\omega ). \end{aligned} \end{aligned}$$
(14)

The ML cost function in the frequency domain is derived analogously to the one in the time domain with the output error covariance matrix \(\varvec{R}\) replaced by the spectral density matrix of the measurement noise. The ML cost function in the frequency domain is, therefore,

$$\begin{aligned} J(\varvec{\varTheta }) = \prod _{i=j}^{n_y} \varvec{\sigma }^2 \left( \varvec{z}_j - \varvec{y}_j \right) \end{aligned}$$
(15)

with

$$\begin{aligned} \begin{aligned}&\varvec{\sigma }^2 \left( \varvec{z}_j - \varvec{y}_j \right) = \frac{1}{N}\sum _{k=0}^{N-1} {\left[ \varvec{z}_j(\omega _k)-\varvec{y}_j(\omega _k) \right] ^*\left[ \varvec{z}_j(\omega _k)-\varvec{y}_j(\omega _k) \right] }, \end{aligned} \end{aligned}$$
(16)

where \((.)^*\) denotes the conjugate transpose of a complex value and \(\varvec{\sigma }^2 (.)\) the model error variance.

Minimization of the cost function from Eq. (11) or Eq. (15) is performed using, e.g., a Gauss–Newton optimization method.

1.2 Frequency response method

The ML method in the frequency domain is based on matching the Fourier transform of the output variables. In contrast, the frequency response method is based on matching the frequency responses, i.e., the ratio of the output per unit of control input as a function of control input frequency.

The frequency response matrix of the identification model \(\varvec{T}(s)\) relates the Laplace transform \(\varvec{Y}(s)\) of the output vector \(\varvec{y}\) to the Laplace transform \(\varvec{U}(s)\) of the input vector \(\varvec{u}\):

$$\begin{aligned} \varvec{Y}(s) = \varvec{T}(s) \varvec{U}(s). \end{aligned}$$
(17)

For the linear state-space system from Eq. (6), the frequency response matrix is determined as

$$\begin{aligned} \varvec{T}(s) = \varvec{C}(s\varvec{I}-\varvec{A})^{-1}\varvec{B} + \varvec{D}, \end{aligned}$$
(18)

where \(\varvec{I}\) denotes the identity matrix.

The quadratic cost function to be minimized for the frequency response method is

$$\begin{aligned} \begin{aligned} J = \frac{20}{N_\omega }&\sum _{k=1}^{N\omega }{ w_\gamma (k) \left[ \left( |T_\mathrm {m}(k)|_{dB}-|T(k)|_{dB}\right) ^2 \right. } \\ &\left. + \; w_{ap} \left( \measuredangle T_\mathrm {m}(k)-\measuredangle T(k) \right) ^2\right] , \end{aligned} \end{aligned}$$
(19)

where T and \(T_\mathrm{{m}}\) are a single-frequency response and its measured counterpart. \(N_\omega\) is the number of frequency points in the frequency interval \([\omega _1,\omega _{N\omega }]\). \(|\ldots |_{dB}\) denotes the amplitude in dB and \(\measuredangle (\ldots )\) the phase angle in degree.

\(w_\gamma\) is an optional weighting function based on the coherence between the input and the output at each frequency. It is defined as

$$\begin{aligned} w_\gamma (k) = \left[ 1.58 (1-e^{\gamma ^2_{xy}(k)}) \right] ^2, \end{aligned}$$
(20)

\(w_{\mathrm{{ap}}}\) is the relative weight between amplitude and phase errors. The normal convention is \(w_{\mathrm{{ap}}} = 0.01745\).

When several frequency responses are approximated together, the overall cost function is the average of the individual cost functions. A good overview of system identification using the frequency response method can be found in Ref. [2].

1.3 PBSIDopt method

The starting point for the PBSIDopt method is a linear discrete-time state-space model in innovation form

$$\begin{aligned} \begin{aligned} \varvec{x}_{k+1}&= \varvec{A}_d \varvec{x}_k + \varvec{B}_d \varvec{u}_k + \varvec{K} \varvec{e}_k \\ \varvec{y}_k&= \varvec{C}_d \varvec{x}_k + \varvec{D}_d \varvec{u}_k + \varvec{e}_k \end{aligned} \end{aligned}$$
(21)

with the input vector \(\varvec{u}_k \in {\mathbb {R}}^{n_u}\), the outputs \(\varvec{y}_k \in {\mathbb {R}}^{n_y}\) and the states \(\varvec{x}_k \in {\mathbb {R}}^{n}\). The innovations \(\varvec{e}_k \in {\mathbb {R}}^{n_y}\) are assumed to be zero-mean white process noise. A finite set of data points \(\varvec{u}_k\) and \(\varvec{y}_k\) with \(k = 1\; \ldots N\) is considered for system identification.

Assuming there is no direct feedthrough, i.e., \(\varvec{D}_\mathrm{{d}}=\varvec{0}\), the system in Eq. (21) is transformed to the one-step-ahead predictor form

$$\begin{aligned} \begin{aligned} \varvec{x}_{k+1}&= \varvec{A}_K \varvec{x}_k + \varvec{B}_K \varvec{z}_k \\ \varvec{y}_k&= \varvec{C}_\mathrm{{d}} \varvec{x}_k + \varvec{e}_k \end{aligned} \end{aligned}$$
(22)

with \(\varvec{A}_K = \varvec{A}_\mathrm{{d}} - \varvec{K} \varvec{C}_\mathrm{{d}}\), \(\varvec{B}_K = \left( \varvec{B}_\mathrm{{d}} \ \ \varvec{K} \right)\) and \(\varvec{z}_k = (\varvec{u}_k \ \ \varvec{y}_k )^T\). Furthermore, it is assumed that all eigenvalues of \(\varvec{A}_K\) are inside the unit circle. Accordingly, the given predictor model is stable. The (\(k+p\))th state \(\varvec{x}_{k+p}\) is given by

$$\begin{aligned} \begin{aligned}&\varvec{x}_{k+p} = \varvec{A}_K \varvec{x}_{k+p-1} + \varvec{B}_K \varvec{z}_{k+p-1} \\&= \varvec{A}_K^p \varvec{x}_k + \underbrace{\begin{pmatrix} \varvec{A}_K^{p-1} \varvec{B}_K&\varvec{A}_K^{p-2} \varvec{B}_K&\ldots&\varvec{B}_K \end{pmatrix} }_{\varvec{\mathcal {K}}^{(p)}} \begin{pmatrix} \varvec{z}_k \\ \varvec{z}_{k+1} \\ \vdots \\ \varvec{z}_{k+p-1} \end{pmatrix} \end{aligned} \end{aligned}$$
(23)

and the (\(k+p\))th output \(\varvec{y}_{k+p}\) is determined

$$\begin{aligned} \varvec{y}_{k+p} = \varvec{C}_\mathrm{{d}} \varvec{A}_K^p \varvec{x}_k + \varvec{C}_\mathrm{{d}} \varvec{\mathcal {K}}^{(p)} \begin{pmatrix} \varvec{z}_k \\ \varvec{z}_{k+1} \\ \vdots \\ \varvec{z}_{k+p-1} \end{pmatrix} + \varvec{e}_{k+p} \end{aligned}$$
(24)

with the extended controllability matrix \(\varvec{\mathcal {K}}^{(p)}\) and the past window length p. Since \(\varvec{A}_K\) is stable, the expression \(\varvec{A}_K^p\) in Eqs. (23) and  (24) can be neglected for large p: \(\varvec{A}_K^p \simeq \varvec{0}\). Therefore, repeating Eqs. (23) and (24) for the (p+1)th in the Nth element yields

$$\begin{aligned} \varvec{X}_{(p+1,N)}&= \varvec{\mathcal {K}}^{(p)} \varvec{Z}_{(1,N-p),p} \end{aligned}$$
(25a)
$$\begin{aligned} \varvec{Y}_{(p+1,N)}&= \varvec{C}_\mathrm{{d}} \varvec{\mathcal {K}}^{(p)} \varvec{Z}_{(1,N-p),p} + \varvec{E}_{(p+1,N)} \end{aligned}$$
(25b)

with

$$\begin{aligned} \varvec{X}_{(p+1,N)} = \begin{pmatrix} \varvec{x}_{p+1}&\varvec{x}_{p+2}&\ldots&\varvec{x}_{N} \end{pmatrix} \end{aligned}$$
(26)

and analogous definitions for \(\varvec{Y}_{(p+1,N)}\) and \(\varvec{E}_{(p+1,N)}\). The merged inputs and outputs are combined as

$$\begin{aligned} \varvec{Z}_{(1,N-p),p} = \begin{pmatrix} \varvec{z}_{1} &{} \varvec{z}_{2} &{} \ldots &{} \varvec{z}_{N-p} \\ \varvec{z}_{2} &{} \varvec{z}_{3} &{} \ldots &{} \varvec{z}_{N-p+1} \\ \vdots &{} \vdots &{} \ldots &{} \vdots \\ \varvec{z}_{p} &{} \varvec{z}_{p+1} &{} \ldots &{} \varvec{z}_{N-1} \end{pmatrix}. \end{aligned}$$
(27)

The predictor Markov parameters \(\varvec{C}_\mathrm{{d}} \varvec{\mathcal {K}}^{(p)}\) are estimated in a least-squares sense with Tikhonov regularization to prevent ill-posed problems. The regularized least-squares problem is given by

$$\begin{aligned} \begin{aligned} \min _{\varvec{C}_\mathrm{{d}} \varvec{\mathcal {K}}^{(p)}}&\left( \left\| \varvec{Y}_{(p+1,N)} - \varvec{C}_\mathrm{{d}} \varvec{\mathcal {K}}^{(p)} \varvec{Z}_{(1,N-p),p} \right\| _F^2 \right. + \left. \lambda ^2 \left\| \varvec{C}_\mathrm{{d}} \varvec{\mathcal {K}}^{(p)} \right\| _F^2 \right) . \end{aligned} \end{aligned}$$
(28)

The regularization parameter \(\lambda\) is chosen with the strong robust generalized cross-validation method, see Ref. [23] for an introduction and a comparison of parameter choice methods.

The estimated predictor Markov parameters \(\varvec{C}_\mathrm{{d}} \varvec{\mathcal {K}}^{(p)}\) can be interpreted as a high-order vector-ARX model (AutoRegressive model with eXogenous input). High-order ARX models based on Eq. (25b) are asymptotically unbiased by correlation issues for large N and p, see Ref. [24]. Thus, this step is essential for subspace identification methods such as PBSIDopt to provide consistent estimates even in correlated closed-loop experiments.

Defining the extended observability matrix \(\varvec{\mathcal {O}}^{(f)}\) with the future window length f

$$\begin{aligned} \varvec{\mathcal {O}}^{(f)} = \begin{pmatrix} \varvec{C}_\mathrm{{d}} \\ \varvec{C}_\mathrm{{d}} \varvec{A}_K \\ \vdots \\ \varvec{C}_\mathrm{{d}} \varvec{A}_K^{f-1} \end{pmatrix}, \end{aligned}$$
(29)

the product of the extended observability matrix \(\varvec{\mathcal {O}}^{(f)}\) and the extended controllability matrix \(\varvec{\mathcal {K}}^{(p)}\) is set up using the estimated predictor Markov parameters \(\varvec{C}_\mathrm{{d}} \varvec{\mathcal {K}}^{(p)}\)

$$\begin{aligned} \begin{aligned}&\varvec{\mathcal {O}}^{(f)} \varvec{\mathcal {K}}^{(p)} \simeq \begin{pmatrix} \varvec{C}_\mathrm{{d}} \varvec{A}_K^{p-1} \varvec{B}_K &{} \varvec{C}_\mathrm{{d}} \varvec{A}_K^{p-2} \varvec{B}_K &{} \ldots &{} \varvec{C}_\mathrm{{d}} \varvec{B}_K \\ \varvec{0} &{} \varvec{C}_\mathrm{{d}} \varvec{A}_K^{p-1} \varvec{B}_K &{} \ldots &{} \varvec{C}_\mathrm{{d}} \varvec{A}_K \varvec{B}_K \\ \vdots &{} \ddots &{} \ddots &{} \vdots \\ \varvec{0} &{} &{} &{} \varvec{C}_\mathrm{{d}} \varvec{A}_K^{f-1} \varvec{B}_K. \end{pmatrix} \end{aligned} \end{aligned}$$
(30)

According to Eq. (25a)

$$\begin{aligned} \begin{aligned} \varvec{\mathcal {O}}^{(f)} \varvec{X}_{(p+1,N)}&= \varvec{\mathcal {O}}^{(f)} \varvec{\mathcal {K}}^{(p)} \varvec{Z}_{(1,N-p),p} \\&= \varvec{\mathcal {U}} \varvec{\mathcal {S}} \varvec{\mathcal {V}}^T \\&= \begin{pmatrix} \varvec{\mathcal {U}}_n&\varvec{\mathcal {U}}_{{\overline{n}}} \end{pmatrix} \begin{pmatrix} \varvec{\mathcal {S}}_n &{} \varvec{0} \\ \varvec{0} &{} \varvec{\mathcal {S}}_{{\overline{n}}} \end{pmatrix} \begin{pmatrix} \varvec{\mathcal {V}}^T_n \\ \varvec{\mathcal {V}}^T_{{\overline{n}}}, \end{pmatrix} \end{aligned} \end{aligned}$$
(31)

the singular value decomposition is applied to reconstruct an estimation of the system states

$$\begin{aligned} \widetilde{\varvec{X}}_{(p+1,N)} = \varvec{\mathcal {S}}_n^{\frac{1}{2}} \varvec{\mathcal {V}}_n^T. \end{aligned}$$
(32)

The model order n corresponds to the n largest singular values in \(\varvec{\mathcal {S}}_n\) used for the state sequence reconstruction.

Finally, the system matrices \(\varvec{A}_\mathrm{{d}}\), \(\varvec{B}_\mathrm{{d}}\), \(\varvec{C}_\mathrm{{d}}\) and \(\varvec{K}\) from Eq. (21) are calculated. First,

$$\begin{aligned} \begin{pmatrix} \widetilde{\varvec{X}}_{(\mathtt {p}+2,N)} \\ \varvec{Y}_{(\mathtt {p}+1,N-1)} \end{pmatrix} = \begin{pmatrix} \varvec{A}_\mathrm{{d}} &{} \varvec{B}_\mathrm{{d}} \\ \varvec{C}_\mathrm{{d}} &{} \varvec{0} \end{pmatrix} \begin{pmatrix} \widetilde{\varvec{X}}_{(\mathtt {p}+1,N-1)} \\ \varvec{U}_{(\mathtt {p}+1,N-1)} \end{pmatrix} \end{aligned}$$
(33)

is solved for \(\varvec{A}_\mathrm{{d}}\), \(\varvec{B}_\mathrm{{d}}\) and \(\varvec{C}_\mathrm{{d}}\) in a least-squares sense. The Kalman gain \(\varvec{K}\) is then calculated from the covariance matrix of the least-squares residuals and the system matrices \(\varvec{A}_\mathrm{{d}}\) and \(\varvec{C}_\mathrm{{d}}\) by solving the stabilizing solution of the corresponding discrete-time algebraic Riccati equation, see Ref. [24] and the references therein.

The inverse bilinear (or any other discrete time to continuous time) transform is then applied to calculate the continuous-time state-space model:

$$\begin{aligned} \begin{aligned} \dot{\varvec{x}}(t)&= \varvec{A} \varvec{x}(t) + \varvec{B} \varvec{u}(t) \\ \varvec{y}(t)&= \varvec{C} \varvec{x}(t) . \end{aligned} \end{aligned}$$
(34)

Selecting only the largest n singular values to reconstruct the state sequence in Eq. (32) already corresponds to a model reduction step. If necessary, further model reduction techniques as described in Ref. [12] can be used to adapt a high-order black box model to the frequency range of interest or to reduce its complexity. In the examples presented in this paper, no further model reduction techniques were applied.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Seher-Weiß, S., Wartmann, J. Initial investigation into the complementary use of black box and physics-based techniques in rotorcraft system identification. CEAS Aeronaut J 11, 501–513 (2020). https://doi.org/10.1007/s13272-019-00431-z

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s13272-019-00431-z

Keywords

Navigation